Cu/CdCO3 catalysts for efficient electrochemical CO2 reduction over the wide potential window

Congwen Sun Jinhui Hao Bing Wei Meng Wu Hong Liu Yusong Xiong Bochen Hu Longhua Li Min Chen Weidong Shi

Citation:  Congwen Sun, Jinhui Hao, Bing Wei, Meng Wu, Hong Liu, Yusong Xiong, Bochen Hu, Longhua Li, Min Chen, Weidong Shi. Cu/CdCO3 catalysts for efficient electrochemical CO2 reduction over the wide potential window[J]. Chinese Chemical Letters, 2023, 34(12): 108520. doi: 10.1016/j.cclet.2023.108520 shu

Cu/CdCO3 catalysts for efficient electrochemical CO2 reduction over the wide potential window

English

  • Using renewable energy generated electricity to drive the electrocatalytic reduction of CO2 into valuable chemicals is an effective way to reduce the greenhouse effect and solve the energy crisis [1]. The electrochemical CO2 reduction reaction (CO2RR) involves multiple steps of electron and proton transfer, so multiple products can be produced [2]. Among the various products of the electrochemical CO2RR, CO is favored because it is an important ingredient in many chemical reactions [3]. Hydrogen evolution reaction (HER) is a strong competitive reaction of CO2RR [4], especially under the relatively negative potential. At a more negative potential, due to the limitation of CO2 mass transfer [5], hydrogen (H2) is more easily generated, so highly selective electrochemical CO2 reduction can only take place in a narrow potential window. Therefore, it is of great significance to develop a catalyst that can inhibit hydrogen (H2) evolution at a more negative potential for realizing the electrochemical reduction of CO2 to CO under a wide potential window.

    Transition metal Cd-based catalysts have been widely studied due to their low cost and good performance in inhibiting H2 evolution [6,7]. CdCO3 in the Cd-based catalyst had the ability to activate CO2 because it contained C atoms, which could combine with O in CO2 [7]. Nevertheless, the activity of CdCO3 for CO2 reduction to CO was not satisfactory, which could only be greater than 90% under a narrow potential window (< 300 mV). Experimental and theoretical studies showed that the construction of heterostructure was one of the key strategies to improve the selectivity of CO2RR products [8]. Heterostructure interfaces with abundant active sites accelerated the electron transport and regulated the adsorption barrier of intermediates [9,10]. For instance, the recent research indicated CdCO3/Cd-CP heterostructure interfaces reduced the energy barrier of the *COOH intermediate, thus realizing the highly selective reduction of CO2 to CO under a 500 mV wide potential window [10]. Although many efforts have been made, the highly selective (> 90%) electrochemical reduction of CO2 to CO over CdCO3 catalysts at a potential window was greater than 500 mV has not been reported.

    Recently, the microenvironment near the electrode has been proved to have a great influence on the selectivity of CO2RR [11]. Raising the local pH near the electrode was of great benefit to inhibit HER, thus improving the selectivity of CO2RR [12,13]. Luo et al. suppressed the formation of H2 by using the OH formed in-situ preparation of rhombohedral CdCO3 crystal (i-CdCO3), and realized the highly selective reduction of CO2 to CO under 500 mV wide potential window [13]. However, the high current density of CO2RR would lead to the OH concentration near the electrode being many orders of magnitude higher than the bulk solution [14,15]. Excessive concentration of OH would react with CO2 to generate CO32−, which was not conducive to inhibition of HER [16]. Therefore, adjusting the concentration of OH near the electrode to maintain an appropriate local pH value was particularly important for inhibiting HER and further improved the selectivity of the catalyst for CO2 reduction.

    Here, we have enhanced the potential window of highly selective reduction of CO2 to CO by combining the construction of heterostructure with the regulation of local pH. By adjusting the feed ratio of Cu2+ and Cd2+, the ability of the electrode to produce OH and adsorb OH could be adjusted. When the ratio of Cu2+/Cd2+ metal ions was 6:4, the catalyst exhibited the FEco exceeded 90% under the wide potential window of 700 mV and the highest FEco of 97.9% at −0.90 V (vs. RHE). Moreover, Cu0.6/CdCO3 showed excellent hydrogen evolution inhibition performance. The experimental results showed that an appropriate high local pH near the electrode inhibited the HER, and Cux/CdCO3 composite structure created more active sites for electrochemical reduction of CO2 to CO. The density functional theory (DFT) calculations also revealed that the excellent reduction of CO2 to CO came from the adjustment of the local electronic structure of adjacent metal active sites by the Cux/CdCO3 composite structure, which reduced the formation barrier of *CO intermediates.

    The synthesis of the catalyst was shown in Fig. 1a. Firstly, Cux/Cd(OH)2 was prepared directly on carbon paper by using a one-step co-deposition method [17], where x represented the molar mass of Cu as a percentage of the total metal ion. Next, the Cux/CdCO3 working electrode was prepared by cyclic voltammetry (CV) conversion in CO2-saturated 0.1 mol/L KHCO3. The microstructure of Cu0.6/Cd(OH)2 and Cu0.6/CdCO3 could be obtained from scanning electron microscope (SEM) and transmission electron microscopy (TEM) measurements (Figs. 1b-d). It could be seen that after the CV treatment, the morphology changed from nanosheets to irregular particles. X-ray diffraction (XRD) directly proved the transition from Cu0.6/Cd(OH)2 to Cu0.6/CdCO3, indicating the successful preparation of the composite catalyst (Fig. 1e). High-resolution TEM images (Fig. 1f) showed interfacial distances of 0.293 and 0.203 nm, which were consistent with (104) plane of CdCO3 [18] and (111) plane of Cu [19], respectively. The element mapping image also showed that Cu, Cd, O, and C species were evenly distributed on the carbon paper (Fig. 1g). The measurement of lattice fringes and the uniform distribution of elements further indicated the successful preparation of a composite catalyst.

    Figure 1

    Figure 1.  (a) Schematic diagram of preparing Cux/CdCO3 electrocatalysts. SEM image of the (b) Cu0.6/Cd(OH)2 and (c) Cu0.6/CdCO3. (d) TEM image of the Cu0.6/CdCO3. (e) XRD patterns of the Cu0.6/Cd(OH)2 and Cu0.6/CdCO3. (f) HRTEM image and (g) EDS-mapping images of the Cu0.6/CdCO3.

    Fourier transform infrared (FT-IR) spectra of the samples before and after CV treatment was used to further demonstrate the successful carbonate transition (Fig. 2a). The disappearance of OH stretching vibration peak and the appearance of anti-symmetric stretching vibration peak and off-plane bending vibration absorption peak of CO32− further proved the successful preparation of Cu0.6/CdCO3 [20,21]. The valence states of Cu and Cd on Cux/CdCO3 surfaces were characterized by X-ray photoelectron spectroscopy (XPS). Compared with Cu0.6/CdCO3, the double peaks of Cd 3d in the CdCO3 had a larger binding energy shift of 0.5 eV (Fig. 2b). On the contrary, the Cu 2p binding energy in Cu shifted to a smaller direction than that in Cu0.6/CdCO3 (Fig. 2c) [22]. XPS results demonstrated that there was a strong interaction between the electronic structures of Cu and Cd [23]. In addition, CO2 adsorption capacity (Fig. 2d) tests were used to analyze the structural properties of the materials. The adsorption capacity of Cu0.6/CdCO3 for CO2 was 2.45 cm3/g STP and 1.70 times that of CdCO3 (1.44 cm3/g STP) (Fig. 2d). Stronger CO2 adsorption capacity was conducive to subsequent CO2 reduction [24].

    Figure 2

    Figure 2.  (a) FT-IR spectra of the Cd(OH)2, CdCO3, Cu0.6/Cd(OH)2 and Cu0.6/CdCO3. (b) Cd 3d XPS spectra of CdCO3 and Cu0.6/CdCO3. (c) Cu 2p XPS spectra of Cu and Cu0.6/CdCO3. (d) CO2 adsorption curves of Cu0.6/CdCO3 and CdCO3 at 273.15 K.

    Fig. 3a showed the linear sweep voltammetry curves (LSV) of Cu0.6/CdCO3, CdCO3, and Cu in Ar and CO2-saturated 0.1 mol/L KHCO3 solutions, respectively. The current density in CO2-saturated 0.1 mol/L KHCO3 solutions was higher than that in Ar-saturated 0.1 mol/L KHCO3 solutions, which preliminarily showed that the catalyst could electrochemically reduce of CO2. According to researches, Cu was a post transition metal with low resistance and good conductivity [25]. CdCO3, which belonged to the metal carbonate class, had a large resistance and a slow electron transfer rate compared to metals [26]. Therefore, at the same potential, CdCO3 had the lowest current density in CO2-saturated 0.1 mol/L KHCO3 solution, while Cu exhibited the highest current density. The current density of catalysts with different proportions gradually increased with the increase of Cu content (Fig. S5 in Supporting information). The performance of the catalyst for CO2 reduction under different applied potentials was tested by constant potential method. The Faraday efficiency (FE) of CO for Cu0.6/CdCO3 exceeded 90% in potential windows ranging from −0.70 V to −1.40 V (vs. RHE) and reached a maximum value of 97.9% at −0.90 V (vs. RHE) (Fig. 3b). By testing the CO selectivity of other samples with different proportions, it was found that the CO selectivity first increased and then decreased with the increase of Cu content, and reached the highest value when the Cu content was 60% (Fig. S7 in Supporting information).

    Figure 3

    Figure 3.  (a) LSV of the Cu, CdCO3 and Cu0.6/CdCO3. (b) FEco of the CdCO3, Cu and Cu0.6/CdCO3. (c) Comparison of CO fractional current densities of CdCO3, Cu and Cu0.6/CdCO3. (d) Stability test of Cu0.6/CdCO3 at −1.20 V vs. RHE.

    In addition to FE, current density, energy conversion efficiency, and stability were also important indicators to evaluate the catalytic performance of CO2RR. Cu0.6/CdCO3 had a superior CO component current density, which could reach a large value of 12.98 mA/cm2 at −1.50 V (vs. RHE) potential (Fig. 3c). This was far more than the CO component current density of CdCO3 (2.98 mA/cm2) and Cu (1.21 mA/cm2), indicating that it had an excellent catalytic activity of CO2 reduction to CO. Moreover, according to the current density and potential function diagram generated by CO (Fig. 3c), when Cu0.6/CdCO3 generated CO, the overpotential required had a positive displacement of about 400 mV compared with that required by CdCO3. These results showed that the composite structures could both increase the current density and reduce the overpotential of electrocatalytic CO2 reduction to produce CO. Furthermore, Cu0.6/CdCO3 had outstanding cathode energy efficiency, which could reach 67.8% at −0.70 V (Fig. S8b in Supporting information).

    Long-term experiments were used to evaluate its stability. The electrodes were tested at a potential of −1.20 V vs. RHE (8.56 mA/cm2) for up to 12 h. As shown in Fig. 3d, the current density dropped slightly during the 12-h test, but the Faraday efficiency was still greater than 90%. The SEM (Fig. S13 in Supporting information), XRD (Fig. S14 in Supporting information), and XPS (Fig. S15 in Supporting information) of the samples did not change significantly before and after the long-term stability test. Therefore, Cu0.6/CdCO3 was a promising catalyst with excellent activity, high CO selectivity, and long-term stability.

    To better understand the source of the enhanced catalyst activity, the electrochemically active surface area (ECSA) was analyzed by testing the double-layer capacitance (Cdl) [27]. The Cdl value of Cu0.6/CdCO3 was 9.23 mF/cm2 (Fig. 4a), 18 and 1.56 times higher than that of CdCO3 (0.51 mF/cm2), and Cu (5.90 mF/cm2), respectively, indicating its abundant active sites on the surface. This result was consistent with the activity of CO2RR. Furthermore, an electrochemical impedance spectroscopy (EIS) test was performed at the potential of −0.90 V (vs. RHE) to analyze the electron transfer resistance between the catalyst and the electrolyte [28]. Nyquist diagram showed that the charge transfer resistance and material transfer resistance of Cu0.6/CdCO3 were much smaller than that of CdCO3 (Fig. 4b). Less resistance was beneficial to the rapid transfer of electrons and the substances [29,30], which could further promote the formation of intermediates in CO2RR. The Tafel curve measurement (Fig. S10 in Supporting information) was used to study the dynamics of the CO2RR process. Compared to Cu (181 mV/dec), CdCO3 (259 mV/dec), and other ratios of Cux/CdCO3 catalysts, Cu0.6/CdCO3 had the smallest Tafel slope (134 mV/dec). The Tafel slope of Cu0.6/CdCO3 was close to 118 mV/dec, indicating that the first electron transfer to CO2 to generate CO2 was a rate-determining step (RDS) [31]. The minimum Tafel slope also meant faster CO formation kinetics [32].

    Figure 4

    Figure 4.  (a) Current density differences (Δj/2) plotted against scan rates of the Cu, CdCO3 and Cu0.6/CdCO3. (b) Nyquist plots. (c) Oxidative LSV curves in Ar-saturated 0.1 mol/L KOH electrolyte of Cux/CdCO3. (d) pH-dependence experiment on Cu0.6/CdCO3 (solid line) and CdCO3 (dotted line) in 0.05 mol/L K2HPO4, 0.1 mol/L KHCO3, 0.1 mol/L KCl solution at 2.71 mA/cm2. (e) In-situ Raman diagram of Cu0.6/CdCO3 at different potentials. (f) Free energy diagrams of each intermediate during the electrochemical reduction of CO2 to CO on the Cux/CdCO3 and CdCO3.

    A rotating disk electrode was used to directly measured the OH generated near the electrode [33]. As the applied potential on the disk increases, CO oxidation peak (Eobs) moved in a more negative direction (Fig. S12 in Supporting information), which directly proved the generation of OH on the electrode. According to the LSV of samples with different proportions, the current density of the sample increased with the increase of Cu content, which meant that the ability to produce OH was also proportional to the Cu content. At the same time, the oxidation test curve in Ar-saturated 0.1 mol/L KOH solution was used to test the overpotential required for OH adsorption (Fig. 4c). The results showed that the overpotential of OH adsorption decreased with the increase of Cu content, which meant that the introduction of Cu could enhance the adsorption of OH [34]. Further comparing the selectivity of different catalysts for H2, it was found that Cu0.6/CdCO3 had the lowest selectivity for H2 over the entire tested potential range (Fig. S7 in Supporting information). Therefore, the pH value of Cu0.6/CdCO3 surface was optimal, which was more conducive to inhibition of HER. To further study the effect of local pH on CO2RR, the CO2RR of Cu0.6/CdCO3 and CdCO3 were tested at a constant current density of 2.71 mA/cm2 at 0.05 mol/L K2HPO4, 0.1 mol/L KHCO3 and 0.1 mol/L KCl, respectively [12]. The buffering capacity of K2HPO4, KHCO3, and KCl decreased in turn, while the local pH value increased in turn. As shown in Fig. 4d, with the increase of local pH, the FEh2 became lower, and the FEco became higher. This further proved that moderately high local pH was favorable for CO2RR and unfavorable for HER.

    In-situ Raman spectroscopy (Fig. 4e) was used to test the reaction path of electrochemical reduction of CO2 to CO by Cu0.6/CdCO3. The peak at 250 cm−1 was attributed to the Cu-CO stretching mode [35], which became stronger with the increase of potential. The peaks at 523 and 1030 cm−1 were attributed to CuxO [35] and *OCO intermediate [36]. After protonation, the *OCO intermediate was transformed into *OCHO, which was conducive to the formation of HCOOH [36]. According to the in-situ Raman results, the peak of *OCO was obvious at high potential (< −1.20 V vs. RHE), indicating that a small amount of HCOOH was generated at high potential, which was consistent with our experimental results. The Cu-OH peak at 706 cm−1 appeared from the −0.70 V vs. RHE [37]. When the applied potential was less than −1.20 V vs. RHE, the peak became more obvious. This directly proved the existence of OH on the surface of the catalyst and further validated the local pH rise. The DFT calculation was performed to further analyze the active source of Cux/CdCO3 for high selectivity CO2RR. The calculated Gibbs free energy (ΔG) diagram of each intermediate was shown in Fig. 4f. On the surface of Cux/CdCO3, the ΔG required for the transition from *COOH to *CO was 1.07 eV, which was much smaller than the 2.97 eV of CdCO3. This indicated that Cux/CdCO3 was more conducive to the formation of *CO [38]. The d-band center was the key to determining the electrochemical activity, which could be evaluated by partial state density (PDOS) data [39]. Compared with CdCO3, the d-band center of Cd in Cux/CdCO3-*COOH was closer to the Fermi level, while the d-band center of Cd in Cux/CdCO3-*CO was further away from the Fermi level (Fig. S19 in Supporting information). This proved that the Cux/CdCO3 surface had a strong adsorption energy for *COOH and a weak adsorption energy for *CO [40]. These results showed that the strong electron interaction between Cu and CdCO3 could reduce the formation energy barrier of *CO intermediate and promote the desorption of CO from Cux/CdCO3 surface, so that the electrochemical CO2 reduction to CO had a fast reaction kinetics.

    In summary, Cux/CdCO3 catalysts were prepared for the first time by simple electrodeposition and CV treatment. Cu0.6/CdCO3 exhibited excellent performance of reducing CO2 to CO, high CO partial current density and fine stability. The experimental results proved that the excellent performance of Cu0.6/CdCO3 catalyst came from the reasonable adjustment of OH generation and adsorption, which inhibited HER. Theoretical calculations have proved that Cux/CdCO3 catalyst reduced the energy barrier from *COOH to *CO intermediate, thus realizing the highly selective electrochemical reduction of CO2 to CO under a wide potential window. This work explored the influence of local microenvironment and electronic structure on the CO2RR, which will provide new guidance for the design of highly selective electrocatalysts under a wide potential window.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was supported by the National Natural Science Foundation of China (Nos. 22225808, 22075111), Sino−German Cooperation Group Project (No. GZ1579), Jiangsu Province Innovation Support Program International Science and Technology Cooperation Project (No. BZ2022045).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2023.108520.


    1. [1]

      S.I. Seneviratne, M.G. Donat, A.J. Pitman, R. Knutti, R.L. Wilby, Nature 529 (2016) 477–483. doi: 10.1038/nature16542

    2. [2]

      K.P. Kuhl, E.R. Cave, D.N. Abram, T.F. Jaramillo, Energy Environ. Sci. 5 (2012) 7050–7059. doi: 10.1039/c2ee21234j

    3. [3]

      X. Chen, W. Zhang, W. Huang, Chin. Chem. Lett. 34 (2023) 107809. doi: 10.1016/j.cclet.2022.107809

    4. [4]

      K. Xu, S. Zheng, Y. Li, et al., Chin. Chem. Lett. 33 (2022) 424–427. doi: 10.1016/j.cclet.2021.07.016

    5. [5]

      D. Raciti, M. Mao, J.H. Park, C. Wang, J. Electrochem. Soc. 165 (2018) F799. doi: 10.1149/2.0521810jes

    6. [6]

      C. Wang, M. Cao, X. Jiang, M. Wang, Y. Shen, Electrochim. Acta 271 (2018) 544–550. doi: 10.1016/j.electacta.2018.03.156

    7. [7]

      X. Jiang, X. Wang, Q. Wang, et al., ACS Appl. Energy Mater. 4 (2021) 2073–2080.

    8. [8]

      J. Sun, W. Zheng, S. Lyu, et al., Chin. Chem. Lett. 31 (2020) 1415–1421. doi: 10.1016/j.cclet.2020.04.031

    9. [9]

      D. Gao, Y. Zhang, Z. Zhou, et al., J. Am. Chem. Soc. 139 (2017) 5652–5655. doi: 10.1021/jacs.7b00102

    10. [10]

      J. Yang, H. Yang, R. Zhang, et al., J. Mater. Chem. A 10 (2022) 23028–23036. doi: 10.1039/D2TA06309C

    11. [11]

      M. Ma, K. Djanashvili, W.A. Smith, Angew. Chem. Int. Ed. 55 (2016) 6680–6684. doi: 10.1002/anie.201601282

    12. [12]

      J.H. Zhou, K. Yuan, L. Zhou, et al., Angew. Chem. Int. Ed. 58 (2019) 14197–14201. doi: 10.1002/anie.201908735

    13. [13]

      J. Xiao, S. Liu, P.F. Sui, et al., Chem. Eng. J. 433 (2022) 133785. doi: 10.1016/j.cej.2021.133785

    14. [14]

      N. Gupta, M. Gattrell, B. MacDougall, J. Appl. Electrochem. 36 (2005) 161–172.

    15. [15]

      B.M. Tackett, D. Raciti, N.W. Brady, N.L. Ritzert, T.P. Moffat, J. Phys. Chem. C 126 (2022) 7456–7467. doi: 10.1021/acs.jpcc.2c01504

    16. [16]

      B.R.W. Pinsent, L. Pearson, F.J.W. Roughton, Trans. Faraday Soc. 52 (1956) 1512–1520. doi: 10.1039/tf9565201512

    17. [17]

      Y. Xiong, B. Wei, M. Wu, et al., J. CO2 Util. 51 (2021) 101621. doi: 10.1016/j.jcou.2021.101621

    18. [18]

      J. Xiao, M.R. Gao, J.L. Luo, J. Power Sources 510 (2021) 230433. doi: 10.1016/j.jpowsour.2021.230433

    19. [19]

      T. Feng, J. Wang, Y. Wang, et al., Chem. Eng. J. 433 (2022) 133495. doi: 10.1016/j.cej.2021.133495

    20. [20]

      A. Askarinejad, A. Morsali, Mater. Lett. 62 (2008) 478–482. doi: 10.1016/j.matlet.2007.05.082

    21. [21]

      L.A. Saghatforoush, S. Sanati, R. Mehdizadeh, M. Hasanzadeh, Superlattices Microstruct 52 (2012) 885–893. doi: 10.1016/j.spmi.2012.07.019

    22. [22]

      S. Kuang, M. Li, X. Chen, et al., Chin. Chem. Lett. 34 (2023) 108013. doi: 10.1016/j.cclet.2022.108013

    23. [23]

      L. Gao, X. Zhong, J. Chen, et al., Chin. Chem. Lett. 34 (2023) 108085. doi: 10.1016/j.cclet.2022.108085

    24. [24]

      G. Si, X. Kong, T. He, et al., Chin. Chem. Lett. 32 (2021) 918–922. doi: 10.1016/j.cclet.2020.07.023

    25. [25]

      Y. Hori, H. Wakebe, T. Tsukamoto, O. Koga, Electrochim. Acta 39 (1994) 1833–1839. doi: 10.1016/0013-4686(94)85172-7

    26. [26]

      R. Zhang, Q. Fu, P. Gao, et al., J. Energy Chem. 70 (2022) 95–120. doi: 10.1016/j.jechem.2022.01.048

    27. [27]

      B. Wei, Y. Xiong, Z. Zhang, et al., Appl. Catal. B 283 (2021) 119646. doi: 10.1016/j.apcatb.2020.119646

    28. [28]

      L. Xu, Z. Wang, X. Chen, et al., Electrochim. Acta 260 (2018) 898–904. doi: 10.1016/j.electacta.2017.12.065

    29. [29]

      J. Zhang, Y. Wang, H. Wang, D. Zhong, T. Lu, Chin. Chem. Lett. 33 (2022) 2065–2068. doi: 10.1016/j.cclet.2021.09.035

    30. [30]

      J. Hao, W. Yang, Z. Peng, et al., ACS Catal. 7 (2017) 4214–4220. doi: 10.1021/acscatal.7b00792

    31. [31]

      S. Zhang, P. Kang, S. Ubnoske, et al., J. Am. Chem. Soc. 136 (2014) 7845–7848. doi: 10.1021/ja5031529

    32. [32]

      S. Liu, H. Tao, L. Zeng, et al., J. Am. Chem. Soc. 139 (2017) 2160–2163. doi: 10.1021/jacs.6b12103

    33. [33]

      F. Zhang, A.C. Co, Angew. Chem. Int. Ed. 59 (2020) 1674–1681. doi: 10.1002/anie.201912637

    34. [34]

      A. Salehi-Khojin, H.R.M. Jhong, B.A. Rosen, et al., J. Phys. Chem. C 117 (2013) 1627–1632. doi: 10.1021/jp310509z

    35. [35]

      T.T.H. Hoang, S. Verma, S. Ma, et al., J. Am. Chem. Soc. 140 (2018) 5791–5797. doi: 10.1021/jacs.8b01868

    36. [36]

      W. Shan, R. Liu, H. Zhao, et al., ACS Nano 14 (2020) 11363–11372. doi: 10.1021/acsnano.0c03534

    37. [37]

      X. Chang, Y. Zhao, B. Xu, ACS Catal. 10 (2020) 13737–13747. doi: 10.1021/acscatal.0c03108

    38. [38]

      H. Xue, H. Zhu, J. Huang, P. Liao, X. Chen, Chin. Chem. Lett. 34 (2023) 107134. doi: 10.1016/j.cclet.2022.01.027

    39. [39]

      W. Yang, J. Li, X. Cui, et al., Chin. Chem. Lett. 32 (2021) 2489–2494. doi: 10.1016/j.cclet.2020.12.057

    40. [40]

      Y. Sun, S. Wang, D. Jiao, et al., Chin. Chem. Lett. 33 (2022) 3987–3992. doi: 10.1016/j.cclet.2021.11.034

  • Figure 1  (a) Schematic diagram of preparing Cux/CdCO3 electrocatalysts. SEM image of the (b) Cu0.6/Cd(OH)2 and (c) Cu0.6/CdCO3. (d) TEM image of the Cu0.6/CdCO3. (e) XRD patterns of the Cu0.6/Cd(OH)2 and Cu0.6/CdCO3. (f) HRTEM image and (g) EDS-mapping images of the Cu0.6/CdCO3.

    Figure 2  (a) FT-IR spectra of the Cd(OH)2, CdCO3, Cu0.6/Cd(OH)2 and Cu0.6/CdCO3. (b) Cd 3d XPS spectra of CdCO3 and Cu0.6/CdCO3. (c) Cu 2p XPS spectra of Cu and Cu0.6/CdCO3. (d) CO2 adsorption curves of Cu0.6/CdCO3 and CdCO3 at 273.15 K.

    Figure 3  (a) LSV of the Cu, CdCO3 and Cu0.6/CdCO3. (b) FEco of the CdCO3, Cu and Cu0.6/CdCO3. (c) Comparison of CO fractional current densities of CdCO3, Cu and Cu0.6/CdCO3. (d) Stability test of Cu0.6/CdCO3 at −1.20 V vs. RHE.

    Figure 4  (a) Current density differences (Δj/2) plotted against scan rates of the Cu, CdCO3 and Cu0.6/CdCO3. (b) Nyquist plots. (c) Oxidative LSV curves in Ar-saturated 0.1 mol/L KOH electrolyte of Cux/CdCO3. (d) pH-dependence experiment on Cu0.6/CdCO3 (solid line) and CdCO3 (dotted line) in 0.05 mol/L K2HPO4, 0.1 mol/L KHCO3, 0.1 mol/L KCl solution at 2.71 mA/cm2. (e) In-situ Raman diagram of Cu0.6/CdCO3 at different potentials. (f) Free energy diagrams of each intermediate during the electrochemical reduction of CO2 to CO on the Cux/CdCO3 and CdCO3.

  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  554
  • HTML全文浏览量:  22
文章相关
  • 发布日期:  2023-12-15
  • 收稿日期:  2023-02-28
  • 接受日期:  2023-04-25
  • 修回日期:  2023-04-23
  • 网络出版日期:  2023-04-28
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章